• Non ci sono risultati.

High-current laser-driven beams of relativistic electrons for high energy density research

N/A
N/A
Protected

Academic year: 2021

Condividi "High-current laser-driven beams of relativistic electrons for high energy density research"

Copied!
16
0
0

Testo completo

(1)

Plasma Physics and Controlled Fusion

PAPER • OPEN ACCESS

High-current laser-driven beams of relativistic electrons for high energy

density research

To cite this article: O N Rosmej et al 2020 Plasma Phys. Control. Fusion 62 115024

(2)

Plasma Phys. Control. Fusion 62 (2020) 115024 (15pp) https://doi.org/10.1088/1361-6587/abb24e

High-current laser-driven beams of

relativistic electrons for high energy

density research

O N Rosmej

1,2

, M Gyrdymov

2

, M M Günther

1

, N E Andreev

3,4

, P Tavana

2

,

P Neumayer

1

, S Z¨

ahter

1,2

, N Zahn

2

, V S Popov

3,4

, N G Borisenko

5

, A Kantsyrev

6

,

A Skobliakov

6

, V Panyushkin

6

, A Bogdanov

6

, F Consoli

7

, X F Shen

8

and A Pukhov

8

1 GSI Helmholtzzentrum für Schwerionenforschung GmbH, Planckstr.1, 64291, Darmstadt, Germany 2Goethe University Frankfurt, Max-von-Laue-Str. 1, 60438, Frankfurt am Main, Germany

3Joint Institute for High Temperatures, RAS, Izhorskaya st.13, Bldg. 2, 125412, Moscow, Russia 4Moscow Institute of Physics and Technology (State University), Institutskiy Pereulok 9, 141700,

Dolgoprudny Moscow Region, Russia

5P. N. Lebedev Physical Institute, RAS, Leninsky Prospekt 53, 119991, Moscow, Russia

6Institute for Theoretical and Experimental Physics named by A. I. Alikhanov of NRC ‘Kurchatov

Institute’, B. Cheremuschkinskaya 25, 117218, Moscow, Russia

7ENEA. Fusion and Nuclear Safety Department, C.R. Frascati, Via Enrico Fermi, 45, 00044, Frascati,

Italy

8Heinrich-Heine-University Düsseldorf, Universit¨atsstraße 1, Düsseldorf, Germany

E-mail:o.rosmej@gsi.de

Received 2 July 2020, revised 13 August 2020 Accepted for publication 25 August 2020 Published 12 October 2020

Abstract

We report on enhanced laser driven electron beam generation in the multi MeV energy range that promises a tremendous increase of the diagnostic potential of high energy sub-PW and PW-class laser systems. In the experiment, an intense sub-picosecond laser pulse of

∼1019Wcm−2intensity propagates through a plasma of near critical electron density (NCD) and drives the direct laser acceleration (DLA) of plasma electrons. Low-density polymer foams were used for the production of hydrodynamically stable long-scale NCD-plasmas.

Measurements show that relativistic electrons generated in the DLA-process propagate within a half angle of 12± 1◦to the laser axis. Inside this divergence cone, an effective electron temperature of 10–13 MeV and a maximum of the electron energy of 100 MeV were reached. The high laser energy conversion efficiency into electrons with energies above 2 MeV achieved 23% with a total charge approaching 1 µC. For application purposes, we used the nuclear activation method to characterize the MeV bremsstrahlung spectrum produced in the interaction of the high-current relativistic electrons with high-Z samples and measured top yields of gamma-driven nuclear reactions. The optimization of the high-Z target geometry predicts an ultra-high MeV photon number of∼1012per shot at moderate relativistic laser intensity of 1019Wcm−2. A good agreement between the experimental data and the results of the 3D-PIC and GEANT4-simulations was demonstrated.

Original content from this work may be used under the terms of theCreative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

(3)

Plasma Phys. Control. Fusion 62 (2020) 115024 O N Rosmej et al

Keywords: relativistically intense laser pulses, near critical electron density plasmas, long-scale plasma channel, direct laser acceleration, low-density polymer aerogels, super-ponderomotive electrons, nuclear reaction yields

(Some figures may appear in colour only in the online journal)

1. Introduction

Laser-driven relativistic electron beams are excellent tools for the generation of ultrashort MeV- gammas [1–4], electron-positron pairs [5–7], neutrons [8], THz [9–11] and betatron [12–17] radiation. In the case of well-directed high current beams of relativistic electrons one can reach extreme high brightness of gamma and neutron sources and use them for radiographic applications in probing of high energy density matter [18,19]. Another application can be found in the field of laser driven nuclear physics [20–22].

Two mechanisms of laser-driven acceleration are currently being discussed as promising for the generation of high energy electrons in near- and sub-critical plasmas. The first one is the laser wake field acceleration (LWFA) [23]: the intense laser pulse drives strong plasma waves that can trap and acceler-ate electrons. The most prominent case of LWFA is the so-called bubble regime [24–27]. The LWFA allows to reach the highest electron energies of 10 GeV [28] and finds its applica-tions in high energy physics and potentially in XFEL devices [29–31]. The LWFA works best in tenuous, very under-dense plasmas and ultra-short laser pulses, shorter than the plasma wavelength.

The second mechanism is the direct laser acceleration (DLA) in a plasma channel created by a relativistic laser pulse [32]. In this case, the electron acceleration occurs in the presence of strong quasi-static electric and magnetic fields generated in plasma [32, 33]. Ponderomotive expulsion of background plasma electrons from the channel caused by a relativistic laser pulse creates a radial electrostatic field and at the same time, the current of accelerated electrons generates the azimuthal magnetic field [33–35]. A relativistic electron trapped in the channel experiences transverse betatron oscil-lations and gains energy efficiently from the laser pulse when the frequency of the betatron oscillations becomes resonant with the Doppler shifted laser frequency [33]. Depending on the plasma density and laser intensity, DLA at the betatron res-onance, stochastic heating [36], transition to wake field accel-eration or a combination of these mechanisms is realized.

The DLA works efficiently in near critical density (NCD) plasmas and for sub-ps laser pulses like petawatt high energy laser for ion experiments (PHELIX) at GSI [37]. Different from LWFA, the DLA does not generate electrons at very high energies, rather, it produces ample amounts of electrons with Boltzmann-like distributions carrying mega-ampere cur-rents. The effective temperature of these electrons can reach several tens of MeV. Interaction of these relativistic electrons with high Z materials causes MeV gamma-radiation that can drive nuclear reactions resulting in neutron production [8].

This scheme based on the laser accelerated electron beam is one of the important pillars of the laser driven nuclear physics program at ELI-NP [20].

Up to now, only a few experiments were performed to demonstrate the advantages of this DLA mechanism in plas-mas of near critical electron density. The energy-transfer from an ultra-intense laser pulse with intensity of 1020 Wcm−2 to hot electrons in NCD plasmas depending on the pre-plasma scale- length was investigated in [38]. In this experi-ment, a one-dimensional expansion of the plasma with a well-controlled scale length was produced by a separate ns laser pulse. The energy distribution of energetic electrons was not measured but simulated using a 2D particle-in cell (PIC) code. A discovered one order of magnitude variation in the coup-ling efficiency of the laser energy into fast electrons, defined via measurements of the Cu-Kα, was explained by the exist-ence of a density gradient optimum that ensures strong laser pulse self-focusing and channeling processes. In [39], meas-urements of electrons accelerated by a relativistic laser pulse propagating across a mm-long extended under-dense plasma plume with (0.02–0.05) × 1021 cm−3 were reported. The experiment showed a strong increase of the effective temperat-ure and a number of super-ponderomotive electrons caused by the increased length of a relativistic plasma channel. New res-ults on the electron acceleration from an ASE pulse pre-ionized foam layers conducted at the Omega EP-laser were reported in [35]. Foam layers of 250 µm thickness and 3 up to 100 mg cm−3 (ne= (0.9–30)× 1021cm−3) mean

dens-ities were used as targets. The 1 kJ short laser pulse had (5.3 ± 1.8) × 1019 Wcm−2 intensity. An approximation of the high energy tail of the measured electron spectra by a Maxwellian-like function resulted in an effective electron tem-perature, of 6–8 MeV averaged over several shots. Measure-ments in the case of foam layers with high areal densities did not show a visible effect toward an increased electron temper-ature. The drawback of this experiment is that the intensity and the duration of the ASE pre-pulse were not adapted to the variety of the used foam densities.

The production of a hydrodynamically stable NCD-plasma layer remains an important issue. A low density polymer aero-gel [42] is a very prospective material for the creation of sub-mm long NCD-plasmas and efficient electron accelera-tion. In [43], polymer foams of 300–500 µm thickness and 2 mg cm−3 mean volume density were used as targets. In foams, the NCD-plasma was produced by a mechanism of super-sonic ionization [44,45] when a well-defined separate ns-pulse was sent onto the foam target forerunning the relativistic main pulse. The required intensity of the ns-pulse (∼5 × 1013Wcm−2) was estimated according to the reference

(4)

[44] in order to reach the velocity of the super-sonic ioniza-tion wave equal to∼250 µm ns−1. The created plasma had an electron density of ne = 0.7 × 1021 cm−3. The

applic-ation of sub-mm thick low-density foam layers provided a substantial increase of the electron acceleration path in a NCD-plasma compared to the case of freely expanding plas-mas created in the interaction of the ns-laser pulse with solid foils. In experiments described in [39], the interaction of sub-ps laser pulses of the moderate relativistic intensity interacted with standard metallic foils and pre-ionized low-density foam layers was investigated. The effective temperature of super-thermal electrons raised from 1.5–2 MeV, in the case of metal-lic foils, up to 13 MeV for the laser shots onto long-scale NCD-plasmas. The high conversion efficiency in the last case was confirmed by the results of the PIC-simulations [40,43]. In [43] it was shown that up to 80% of the laser energy was converted into electrons, while up to 7% was absorbed by the C, O and H-ions.

In continuation of our research on the electron accelera-tion in long-scale NCD-plasmas described in [43], we present new results on the characterization of the super-ponderomotive electrons and demonstrate a big advantage of our approach for generation of ultra-bright gamma sources compared to the interaction of ultra-high intensity laser pulses with conven-tional foil targets. The experiment was supported by PIC and GEANT4 simulations.

The paper is organized as follows: laser and target paramet-ers together with the used experimental set-up are described in details in section 2; Experimental results on character-ization of super-ponderomotive electron beams are presen-ted in section3; in section4, results of PIC-simulations, that account for the experimental geometry, are compared to the experiment; in section5, the experimental and numerical res-ults on the gamma-ray generation by the well-directed ultra-relativistic electron beams are discussed; section6summarizes the results.

2. Experimental set-up

Experiments were performed at the PHELIX at the Helmholtzzentrum GSI Darmstadt [37] at the highest ns laser contrast≥1011. A s-polarized laser pulse of 1.053 µm funda-mental wavelength delivered by the Nd:glass laser was sent onto targets at 5–7 degree to the target normal. Two different focusing off-axis parabolic mirrors were used providing peak laser intensities of (1–2.5)× 1019 Wcm−2 (aL = 2.7–4.27)

and (7–10)× 1020Wcm−2(aL= 22.6–27.0). Here aL is the

normalized vector potential that scales as aL2= 0.73· IL,18· λ2

with the laser intensity IL,18 normalized to 1018Wcm−2and

the laser wavelength λ, in µm. The duration of the laser pulse was 750± 250 fs. In the case of a moderate relativistic laser intensity (1–2.5) × 1019 Wcm−2, 90 ± 10 J laser energy measured after the main amplifier was focused by means of a 150 cm off-axis parabolic mirror into an elliptical focal spot with FWHM diameters 12± 2 µm and 18 ± 2 µm con-taining a laser energy of EFWHM ≃ (17–22) J. In the case of

the ultra-relativistic laser intensity, 180 J laser energy was focused into a 2.7± 0.2 µm × 3.2 ± 0.2 µm focal spot by a

40 cm off-axis parabolic mirror. The laser energy in the focal spot was 20% of that after the main amplifier and reached

EFWHM≃ (36–40) J. The laser spot size on target and the laser

energy in the focal spot were controlled in every shot. Exper-iments on the direct laser acceleration of electrons in plasmas of near critical density were performed using the mentioned off-axis parabolic mirror with a focal length of 150 cm. In the case of ultra-high laser intensity, shots were done only onto metallic foils.

Cellulose triacetate (TAC, C12H16O8) layers of 2 mg cm−3 volume density and 300–400 µm thickness [42,43] were used as targets [43]. In the case of fully ionized plasma, it cor-responds to 0.64 × 1021 cm−3 electron density or 0.64 n

cr,

where ncr = 1021 cm−3. A sub-mm long NCD-plasma was

produced by sending a well-controlled nanosecond pulse fore-running the relativistic main pulse onto a foam. The intens-ity of the ns laser pulse was kept at∼5 × 1013Wcm−2level in order to initiate a super-sonic ionization wave propagating with 2× 107cm s−1velocity, see [44,45] for more details. The ns-pulse was focused on target by the same parabolic mir-ror as the short pulse. Generation of a long scale NCD-plasma requires focusing optics with hundreds of micrometers long Rayleigh length that was the case for the 150 cm off-axis para-bolic mirror. This ensures rather constant ns laser pulse intens-ity along the whole layer thickness. The delay between the peak of the ns-pulse and the relativistic main pulse was fixed by 2–3 ns. Experimental set-ups and a target holder with a low-density foam layer fixed inside a 2.5 mm in diameter Cu-washer are shown in (figure1). In both set-ups, the electron spectra were measured simultaneously at three different angles with respect to the laser axis. The spectrometers were equipped with 0.99 T static magnets and imaging plates (IP) [48,49] for the detection of the electron signal. The energy resolution of the spectrometers with 300 µm (width)× 1000 µm (height) entrance slits was numerically simulated using 2D B-field dis-tributions measured inside the spectrometer steel-housing. It was shown that an experimental error in the energy measure-ments is caused mostly by the entrance slit and is not higher than 2%. This allows for reliably measuring of electron spec-tra from 1.75 up to 100 MeV. The spectrometers were placed in the horizontal plane XOZ (which is perpendicular to the laser polarization along OY) at a distance of 405 mm from the interaction point around the target at 0, 15and 45(5–7 in figure1(a)). Massive 20 up to 40 mm long WCu—collimator blocks with 3 mm entrance hole were placed in front of every spectrometer in order to shield the front plate from gamma-rays and to increase the signal-to-noise ratio.

The angular distribution of the electron beam in a wide range of angles was measured by means of a stack of three stainless steel cylindrical plates of 3 mm thickness each, rather similar to that described in [50]. The cylinder-stack had a curvature radius of 200 mm and was placed 230 mm away from the target position (4 in figure 1(a)). The observation angle was 0◦± 50◦in the horizontal direction and 0◦± 15◦ in the vertical direction, where 0 corresponds to the laser axis. A horizontal 4 mm wide slit centered at the laser pulse height allowed for the electron propagation to the magnetic spectrometers placed behind the cylinder-stack. In order to

(5)

Plasma Phys. Control. Fusion 62 (2020) 115024 O N Rosmej et al

Figure 1. Experimental set-ups (top-view) used for shots (a) at moderate and (b) ultra-high relativistic intensities. Scheme (a) shows an off-axis focusing parabola (1), laser beam (2), target position (3), stack of cylinders (4), and three magnetic spectrometers (5÷ 7). The laser intensity distributions for both set-ups are also shown.

map the position of the electron beam in space, small holes with a 20 mm interval in vertical and horizontal directions were drilled into the front plate. Large area IPs were placed between the first and the second, and the second and the third cylindrical plates to map the spatial distribution of electrons with E > 3 MeV and E > 7.5 MeV respectively.

A nuclear activation-based diagnostic [51,52] was used to characterize MeV bremsstrahlung radiation generated in the interaction of the super-ponderomotive electrons with high-Z materials. The activation samples, consisting of stacked together Au, Ta, In and Cr plates, were fixed at horizontal angles of 6 ± 1◦ and 16 ± 1◦ to the laser pulse axis (figures2(b) and (d)). After irradiation by electrons and gam-mas, all activation samples were counted multiple times on a low background HPGe-detector to identify nuclides via known γ-ray energies, intensities, and half-life times. Reaction yields of the obtained isotopes 196 194,192Au, 180,178mTa and 51,49Cr produced due to photodisintegration were used for reconstruc-tion of the MeV bremsstrahlung spectrum. A 0.25 mm thin In-foil was placed between Ta and Cr for measurements of the neutron yield [53].

3. Experimental results on high-current relativistic electrons

Figure2shows raw electron spectra measured in two selected laser shots at 0with respect to the laser axis by means of the 0.99 T electron spectrometer and the angular distribution of the electron beam with E > 3 MeV and E > 7.5 MeV meas-ured using the cylinder-stack. For registration of the electron signal Fuji BAS IP MS type was used as detector, while for diagnostic of the electron angular distribution Fuji BAS IP

TR. For each picture in (figure2), the maximum intensity of the related raw signals expressed in PSL (Photo Stimulated Luminescence) is indicated. The presented shots were made at ∼1019 Wcm−2 laser intensity onto a 10 µm thin Au-foil (figures2(a) and (b)) and a pre-ionized foam combined with a 10 µm Au-foil attached to the rear side (figures2(c) and (d)). In the case of the pre-ionized foam, the raw electron spectrum shows a 10-fold enhancement of the IP-signal (compare fig-ures2(a) and (c)) that reflects a corresponding increase of the number of electrons with energies above 2 MeV [48]. In the case of the foil target, the maximum of the measured electron energy lays in the area up to 15 MeV, while for shots onto the pre-ionized foam stacked together with the Au-foil, the max-imum electron energy reached 95–100 MeV.

For the shots onto NCD-plasmas, we observed by means of the cylinder-stack a strong collimation of electrons with energies E > 3 MeV (first IP) and E > 7.5 MeV (second IP) into a well-directed electron beam with a half of a divergence angle of 12◦± 1◦at FWHM (figure2(d)). The IP signal pro-duced by electrons with energies higher than 3 MeV that could pass through the first 3 mm of stainless steel was oversaturated (>300 PSL), while the IP signal obtained for the laser shot onto foil reached only 8–10 PSL (figure2(b)).

GEANT4 [56–58] simulations were performed to define the input of electrons and x-rays into the IP-image. The discussed experimental geometry (figure 1(a)), the measured electron energy distribution (figures 2(c), 5 and 0) and the BAS IP TR imaging plate response to electrons and photons [48] were used as input parameters. Simulations showed that the contribution of photons to the IP signal is less than 5% and that the obtained signal can be attributed mainly to the elec-tron angular distribution.

(6)

Figure 2. Raw electron spectra measured along the laser axis by means of the 0.99 T electron spectrometer (a, c) and angular distribution of the electron beam with E > 3 MeV and > 7.5 MeV registered using the cylinder-stack (b, d): (a, b) shot at 1.6× 1019Wcm2laser intensity onto a 10 µm thin Au-foil; (c, d) shot at 1.5× 1019Wcm2laser intensity onto pre-ionized foam layer of 325 µm thickness combined with a 10 µm Au-foil.

Figure 3. Comparison of the (a) IP image obtained in experiments using the cylinder-stack and (b) GEANT4 simulation result. ‘Needles’ in the 3D-representation of the image are caused by the holes drilled in the first steel cylinder at a distance of 2 cm from each other to map the electron distribution spatially.

The electron energy distribution measured by three spectro-meters placed in the horizontal plane XOZ at 0◦, 15and 45 to the laser axis (figure1(a)) is shown in (figure4). The shot was made at 1.5× 1019Wcm−2laser intensity onto a 325 µm thick foam layer pre-ionized by the ns pulse.

One can see that the majority of electrons is acceler-ated along the laser axis (0). The measured electron spec-tra were approximated by a Maxwellian-like distribution functions with one or two temperatures. The electron energy distribution measured at 0was approximated by one effective temperature T1 = 12.0 ± 1.4 MeV. The temperature and

the number of accelerated electrons drops down to T1 8.0 MeV (T2≃ 11.0 MeV) at 15◦and further to T1≃ 0.9 MeV (T2≃ 2.9 MeV) at 45◦.

For the first time, we explored advantages of a com-bination of foam layers with thin metallic foils in shots at

∼1019 Wcm−2 laser intensity and demonstrated very stable

electron signals up to 100 MeV energies. Spectra measured at 0, 15 and 45 in the interaction of the 1.9 × 1019 Wcm−2laser pulse with the foam layer stacked together with a 10 µm thin Au-foil (figures2(b) and5) show a strong angular

(7)

Plasma Phys. Control. Fusion 62 (2020) 115024 O N Rosmej et al

Figure 4. Energy spectra of electrons per steradian measured at 0(red), 15(green) and 45(blue) to the laser propagation direction for a shot onto the pre-ionized foam layer at 1.9× 1019Wcm2laser intensity.

Figure 5. Electron spectra measured at 0(red), 15(green) and 45(blue) to the laser pulse axis for shots at∼ 1019Wcm2laser intensity. Shots were made onto the pre-ionized foam layer combined with 10 µm Au-foil and directly onto the foil (gray).

dependence similar to that in (figure 4). Additionally, one observes high energy (E > 25 MeV) electron tails at 0and 15 that can be described by an exponential function with a very high second effective temperature of T2= 28.0± 3.4 MeV for 0and T2= 19.1± 2.3 MeV for 15◦.

The length of NCD-plasma plays a crucial role in the electron energy gain. Despite moderate relativistic laser intensity, the effective electron temperature measured in our experiment is three times higher than those measured in the interaction of 2 × 1020 Wcm−2 laser intensity with double

(8)

Figure 6. Electron energy distributions are normalized to the corresponding laser energies inside FWHM of the focal spot in Joules. Shots were made onto the pre-ionized low-density foam layer in combination with 10 µm Au-foil at 1.5× 1019Wcm2laser intensity (red) and onto 10 µm thin Ti-foils irradiated by the 9× 1020Wcm2(green) and 1.6× 1019Wcm2(gray) laser intensities. In all cases, spectra were measured along the laser axis (0).

layered targets, which consist of an NCD slab and an over-dense foil [54]. A major reason for this effect is a long accel-eration path provided by a 300 µm thick foam layer, while the length of the NCD-plasma in [54] was of several µm only.

In contrast to foams, shots onto 10 µm Ti- or Au-foils at ∼1019 Wcm−2 laser intensity shows a near isotropic angular distribution with effective temperatures between 1.4± 0.1 MeV (45◦) and 2.2± 0.2 MeV (0◦) (gray signal in figures5and6).

In this experimental campaign, the electron energy distri-bution was also measured in the interaction of ultra-relativistic laser pulses of 1021Wcm−2intensity (figure1(b)) with metal-lic foils. Figure6shows the electron signals measured at 0 in the shot made onto the pre-ionized low density foam layer in combination with 10 µm Au-foil at 1.5 × 1019 Wcm−2 laser intensity (red) and onto 10 µm thin Ti-foils irradi-ated at 9 × 1020 Wcm−2 (green) and 1.6 × 1019 Wcm−2 (gray) laser intensities. The number of accelerated electrons presented in (figure6) is normalized to the laser energy con-tained in the FWHM of the focal spot, since in the case of the short focusing parabolic mirror this energy was twice higher than in the long-focus case (36–40 J vs. 17–20 J). The effective temperature of the major fraction of electrons in the case of the ultra-relativistic laser intensity is estimated as high as 6.7± 1.2 MeV, which is lower than for shots onto pre-ionized foams at 1019 Wcm−2 (11–12 MeV). Another important difference can be observed analyzing the maximum of the measured electron energy. In contrast to shots with the ultra-relativistic laser intensity, where the maximum of the measured electron energy does not exceed 40 MeV, the

electron spectra, measured at 0 by irradiation of foams in combination with thin foils, show a very stable signal up to 90–100 MeV electron energy (figures5and6).

Once the optimal parameters for the ns-pulse that drives the super-sonic ionization inside the foam were found, the DLA mechanism in the relativistic plasma channel reaches a high level of reproducibility. In measurements at 0, the effective temperature of the main fraction of super-ponderomotive electrons, averaged over 10 shots onto the pre-ionized foam layers only and onto the combination of foams and foils, reaches 11.5± 2.0 MeV.

Figure 7 summarizes data on the number of electrons in different energy ranges and the maximum of the measured electron energy for selected shots made onto various targets irradiated at two laser intensities: (1.5–1.9)× 1019Wcm−2 (blue) and 9.0 × 1020 Wcm−2 (red-dashed) and measured at 0, 15 and 45. Data is normalized to laser energy in the selected shot. The experimental error of ∼25% that occurs by the evaluation of the electron number is caused mostly by the uncertainty of the IP response to the tron impact. This uncertainty remains constant for all elec-tron energies above 0.1 MeV [48]. Additionally, the IP sig-nal fading [49] was taken into account. In the experiment, the IPs were scanned firstly 30–50 min after each laser shot, so that for the signal corrections, a fading factor of 0.65–0.7 was used.

The full height of the bars in (figure7) corresponds to the total amount of electrons with energies E > 2 MeV in steradian normalized to the laser energy contained in the FWHM of the focal spot. For laser intensities of (1–2.5)× 1019Wcm−2this

(9)

Plasma Phys. Control. Fusion 62 (2020) 115024 O N Rosmej et al

Figure 7. Number of electrons Neper steradian and per J laser energy for different electron energy ranges for selected shots onto various

target types irradiated at two laser intensities: 1.5÷ 1.9 × 1019Wcm2(blue) and 9× 1020Wcm2(red, dashed) and measured at 0, 15 and 45◦. The maximum of the measured electron energy Emaxis also shown.

energy was of (17–22) J, while for ultra-relativistic intensity of 1021Wcm−2it reached (36–40) J. The discussed uncertainty in the electron number caused by the IP response is valid in every energy range presented in (figure 7). In the diagram, this uncertainty is shown for the total number of electrons. In the case of the foam target, one counts 7× 1010sr−1J−1 electrons with E > 2 MeV and 4.4 × 1010 sr−1 J−1 elec-trons with E > 7.5 MeV. In the case of the foam layer stacked together with the Au-foil, the total number of electrons with E > 2 MeV is close to the foam-case, while the number of electrons with E > 7.5 MeV is slightly higher and reaches the value of 5.0× 1010sr−1 J−1. The darkest colored part rep-resents electrons with E > 40 MeV with the highest number obtained in the case of the combination of the foam layer with the 10 µm thin Au-foil. This part of the electron spectrum is of interest e.g. in respect to (γ, xn) reactions with tens of MeV threshold gamma-energy.

In results obtained at relativistic laser intensity of 9 × 1020 Wcm−2 in the shot onto Ti-foil this part is miss-ing. Although the role of the thin foils in the combination with foams demands further theoretical and experimental investigations, their positive effect is clearly seen by comparing the values of the measured maximum electron energy. In the cases of ‘foam + Ti’ and ‘foam + Au’ (fig-ure7) they lay between 95 and 100 MeV, instead of 70 MeV for the case of foam only.

4. PIC-simulations

3D PIC simulations of the laser propagation in the NCD plasma were performed using the Virtual Laser Plasma

Laboratory (VLPL) code [55] for the laser parameters and interaction geometry used in the experiment. In particular, the laser pulse intensity in time and space was approximated by a Gaussian distribution. Elliptical form of the focal spot was taken from the experiment with FWHM axes 11 µm in a vertical and 15 µm a horizontal direction. The laser pulse energy in the FWHM focal spot of 17.5 J and the FWHM pulse length of 700 fs resulted into the laser intensity of 2.5× 1019Wcm−2with aL= 4.28. The homogeneous plasma

was composed of electrons and fully ionized ions of car-bon, hydrogen and oxygen. Simulations accounted for the ion type and the ion fraction in accordance with the chemical composition of triacetate cellulose C12H16O8, see e.g. [46, 47]. The simulation box had sizes of 350 × 75 × 75 µm3. The first 10 and the last 15 µm from the total 350 µm of the space in x-direction (the laser axis) were free of the plasma initially. Sizes of a numerical cell were 0.1 µm along the x-axis and 0.5 µm along the y-axis and the z-axis. The number of particles per cell in the simulation was 4 for the electrons and 1 for the ions of each type. Boundary conditions were absorbing for particles and fields in each direction.

The initial electron density (together with the neutraliz-ing ion density) at the moment of the main pulse arrival was of 0.65 ncr with a step-like profile. Previously,

PIC-simulations were performed for a step-like density profile with

ncr and 0.5 ncr [40, 41] and for a partially ramped

dens-ity profile in order to account for plasma expansion toward the main laser pulse [43]. The simulations result in a very similar overall behavior of the energy and angular distribu-tions of super-ponderomotive electrons in all mentioned cases.

(10)

Figure 8. 3D view of the laser intensity distribution at time moments t1= 100 fs (a) and t2= 433 fs (b). The time t = 0 corresponds to the

maximum laser pulse intensity on the target front side (x = 10 µm).

Figure8gives a 3D view of the laser intensity by propagation of the laser pulse in the NCD plasma at time moments t1= 100

fs (a) and t2= 433 fs (b).

The laser pulse propagates from the left to right and at t = 0 its maximum intensity is on the target front side (x = 10 µm). We observe strong self-focusing (a) and later filamentation (b) of the laser pulse that produces the main channel and, at the end of interaction, a few side channels inside the plasma. For the discussed moderate laser energy (∼20 J), the filamentation of the laser pulse takes place only by the end of the accelera-tion process (at t2= 433 fs, figure8(b)). In this case, the main

fraction of accelerated electrons already gained the energy in the free part of the channel and the laser filament-ation weekly effects the energy gain and the angular distri-bution of the electron beam. With an increased laser energy, filamentation starts to be much more pronounced and occurs at earlier times [40]. It manifests itself in the filamentation of the current of accelerated electrons and in the worse collimation of the generated electron beams. Nevertheless, simulations show that the DLA process continues to be very effective [40].

The laser ponderomotive force expels background plasma electrons off the channel and creates a radial quasi-static elec-tric field. At the same time, the current of the accelerated relativistic electrons generates an azimuthal magnetic field. The structure of the quasi-static self-generated E and B-fields was analyzed in detail in [32–35] and in publications that con-sider PHELIX-laser parameters [40,43]. In [43] it is shown that the electron current density in the relativistic plasma channel reaches 1011A cm−2and the quasi-static azimuthal magnetic field exceeds 100 MG.

Figure 9 presents snapshots of the electron phase space 100 fs after the laser pulse peak intensity arrived at the left plasma boundary. Acceleration occurs mostly in the laser pulse propagation direction x, as it can be seen in (figure9(a)), where the electron momentum px grows approaching∼100

mc. The high values of the transversal electron momentum py∼ 30–50 mc (laser polarization direction) (figure9(b)) and

pz∼ 20 mc (figure9(c)) are much larger than the normalized

amplitude of the laser pulse a0= eE0/mcω∼ 5. These large values of the electron transverse momenta are due to the res-onance between the electron betatron frequency in the chan-nel fields and the Doppler shifted laser frequency [32, 33]. Then, magnetic field converts the gained transverse particle

momentum into the longitudinal one via the v x B force. Figure9(d) zooms the pyvs x electron phase space in the range

[43–70] x/λ, where we clearly see that the electron transverse oscillations are modulated at the laser phase.

In the PIC simulation, we registered electrons leaving the NCD plasma. Angular distribution of the electrons with

E > 7.5 MeV is shown in (figure 10) in spherical coordin-ates with a polar axis OX along the laser propagation direc-tion: θ = arctan(

p2

y+ p2z/px), φ = arctan(pz/py). One can

see that a high fraction of the super-ponderomotive electrons is accelerated in the laser direction and propagate in the rather tight divergence cone with a half angle of 10-12. This res-ult supports the experimental measurements presented in (fig-ure3).

The 3D capability of the PIC code allows for simulations close to the real experimental conditions. Thus, the abso-lute energy spectra, i.e. the number of accelerated electrons in any energy range and their angular distribution can be obtained.

The electron spectra presented in (figure 11) were sim-ulated in the horizontal plane for θ = 0◦ ± 2◦ (red dots), 15 ± 2◦(green dots) and 45 ± 2◦ (blue dots), which cor-respond to the positions of the electron spectrometers in the experiment (figure1(a)). The obtained effective temperatures are in good agreement with the experimental results presented in (figure4) for all three observation directions. According to simulations, the total number of electrons with E > 2 MeV, propagating in 2π, reaches the value of 5× 1012that corres-ponds to∼ 1 µC of the electron charge and 27% conversion of the laser energy into relativistic electrons. Current simula-tions show that only 12% of electrons with E > 2 MeV are contained in the 0.16 sr divergence cone. Another situation is with the super-ponderomotive electrons with energies beyond 7.5 MeV that are relevant for gamma-driven nuclear reactions. Simulations result into Ne= 3.3× 1011electrons contained in

0.16 sr, that is 30% of all super-ponderomotive electrons with E > 7.5 MeV propagating in 2π. This number is in good agree-ment with the experiagree-mental value of (2± 0.6) × 1011.

Based on the results of simulations and measurements, one can estimate a charge carried by the super-ponderomotive electrons with E > 2 MeV and E > 7.5 MeV propagating along the laser axis inside the divergence cone of 0.16 sr as high as ∼100 and ∼50 nC with corresponding conversion efficiencies of 12% and 6%.

(11)

Plasma Phys. Control. Fusion 62 (2020) 115024 O N Rosmej et al

Figure 9. Snapshots of the electron phase space 100 fs after the laser pulse peak intensity arrived at the left plasma boundary: (a)

momentum pxvs x; (b) momentum pyvs x; (c) momentum pzvs x; (d) zoomed part of momentum pyvs x in the range x/λ = [43, 70]. Color

panels present a number of pseudo-electrons in simulations.

Figure 10. Angular distributions of electrons dN/d Ω [sr−1] with energies E > 7.5 MeV that left the simulation box by the time

t = 2.5 ps.

5. Gamma-ray generation

High-current and well-directed electron beams with energies of several tens of MeV are excellently suited for the produc-tion of ultra-intense gamma sources by penetraproduc-tion of high Z materials.

In our experiment, the MeV bremsstrahlung spectrum was evaluated from the yields of gamma-induced nuclear reac-tions as it was done in [52]. For this purpose, 1 mm-thick activation samples of Au, Ta, Cr and 0.25 mm thin In-foil of 15× 15 mm2 area were placed at 6 ± 1◦ and 16 ± 1◦ to the laser pulse propagation direction 150–230 mm away from the target position (figures 2(c) and (d)). The shot at 1.5 × 1019 Wcm−2 laser intensity was made onto the pre-ionized foam layer stacked together with 1 mm thick

Au-plate. After irradiation by electrons and gammas, all activation samples were counted multiple times on a low back-ground HPGe-detector to identify nuclides via known γ-ray energies, intensities, and half-life times. High reaction yields of the isotopes 196 194,192Au, 180,178mTa and51,49Cr measured in the activation samples at 5to the laser axis are shown in (figure12). The isotope yields measured in the samples placed at 15 to the laser axis were at least one order of magnitude lower.

The normalization of the experimental data presented in (figure12) was made by taking the laser energy in the FWHM of the laser focal spot (20 J in case of 1019 Wcm−2 and 40 J in the case of 1021Wcm−2laser intensities) and a solid angle of 7× 10−3sr covered by the activation samples into account. For determination of the reaction yield, the effi-ciency of the HPGe-detector, the intensities of γ-lines and the detector dead time were taken into consideration. The uncertainty in the reaction yields is mainly due to the uncer-tainty in the number of counts in the full energy peaks. Dark blue colored bars in (figure 12) show the reaction yields obtained in the shot at 1.5 × 1019 Wcm−2 laser intensity onto a pre-ionized foam layer stacked together with 1 mm thick Au-plate. The reaction yields for 1021 Wcm−2 laser intensity are presented by light sand colored bars. In shots at (1–2) × 1019 Wcm−2 laser intensity, we observe up to 50-times higher 194Au isotopes yields with the threshold gamma energy Eγ ≃ 22 F-23 MeV compared to the laser shots directly onto the convertor plate at 1021 Wcm−2. The 192Au isotopes created in (γ, 5n) reactions with the cross-section threshold at Eγ≃ 38.7 MeV were observed in shots onto pre-ionized foams at ∼1019 Wcm−2 only. The sub-stantial difference in the isotope production yield for these two cases of the laser-matter interaction can be explained by the difference in the measured electron spectra shown in (figures5and6).

(12)

Figure 11. Energy distributions of electrons per steradian (d2N/dEdΩ) that leave the simulation box at t = 2.5 ps for three different angles θ = 0◦(red), 15(green) and 45◦(blue) in the horizontal plane XOZ.

Figure 12. Experimental data on (γ, xn) reaction yields in Au, Ta und Cr and results of GEANT4 simulations normalized to steradian and the laser energy in the corresponding shot. Blue colored bars present reaction yields obtained in the shot at 1.5× 1019Wcm2laser intensity onto a pre-ionized foam layer stacked together with 1 mm thick Au-plate used as a radiator. Green bars show reaction yields simulated for this case with GEANT 4. The reaction yields measured in the shot at 1021Wcm2directly onto the radiator plate is shown by light sand colored bars. In the gray row below the figure, the threshold photon energies for the production of the corresponding isotope are presented.

The bremsstrahlung spectra reconstructed from the meas-ured Au, Ta and Cr isotopes yields were described by an exponential function with an effective temperature of 12.7 ± 2.4 MeV for the laser shot at 1.5 × 1019 Wcm−2 intensity and 5.4 ± 1.7 MeV in the case of the laser shot

at 1021Wcm−2. Interaction of the super-ponderomotive elec-trons with high-Z samples resulted in a gamma fluence Nγ (>7.5 MeV) = 2.0± 0.5 × 109cm−2at 180 mm distance. This is 10-times higher than the value obtained in the laser shot at ultra-relativistic laser intensity.

(13)

Plasma Phys. Control. Fusion 62 (2020) 115024 O N Rosmej et al

Figure 13. Electron and gamma fluencies (E > 7.5 MeV) in arbitrary units in dependence on the penetration depth in the stack of activation samples Au/Ta/In/Cr.

Table 1. Number of MeV-photons Nphand effective temperature Teffof the bremsstrahlung spectrum generated by super-ponderomotive

electrons in 1 cm× 1 cm × 0.5 cm Au-block in dependence on the penetration depth.

0.1 mm 1 mm 3 mm 5 mm

Au-radiator thickness

Energy range Teff, MeV Nph Teff, MeV Nph Teff, MeV Nph Teff, MeV Nph

1–7 MeV 2.5 5× 1010 2.5 2× 1011 2.5 7× 1011 2.5 6× 1011

7–70 MeV 6.1–11.4 6× 1010 5.9–10.7 1× 1011 5.7–9.8 2× 1011 5.6–9.3 1.6× 1011

The modeling of the electron beam interaction with the activation samples was performed by means of the GEANT4 Monte Carlo code [56–58] for the discussed experimental geo-metry. Green bars in (figure12) show simulated reaction yields that are in a good agreement with the experiment. In the mod-eling, the measured electron energy distribution together with the divergence angle of super-ponderomotive electrons were used as input parameters. The activation samples of Ta, Au, In and Cr were ‘placed’ at 5to the laser axis 180 mm away from the laser-foam interaction point. The electron beam traversed the 1 mm Au-radiator producing MeV-bremsstrahlung. The high fraction of the relativistic electrons that escape the radi-ator together with gamma-radiation propagated in vacuum and interacted with the activation samples.

For the simulations, cross sections for (γ, xn) reactions in Au, Ta and Cr were taken from [56]. A cross check between these cross sections and those from [59] used for the reconstruction of the gamma-spectrum from experimentally measured isotope yields provided a good agreement. Figure12 shows the transformation of the relativistic electrons with E > 7.5 MeV, which heat the Au-activation sample from the left side at first, into gamma-rays in dependence on the penet-ration depth in the stack. Vertical cross sections in (figure13) made through the central part of the stack show a normal-ized electron (left panel) and a photon (right panel) fluence. The initial level of the bremsstrahlung radiation of 0.55 a.u. that heats the stack is generated by the super-ponderomotive

electrons in the 1 mm thick Au-radiator, an additional gamma-fluence of ≥0.3 a.u. was produced by the relativistic elec-trons in the activation samples. A slight increase of the electron number in the Au-vacuum interface is caused by gamma-rays.

The simulations result in the bremsstrahlung radiation with effective temperatures of T1≃ 7.1 MeV (Eγ= 10–20 MeV) and T2≃11.2 MeV (Eγ= 25–60 MeV). A peak photon num-ber of 1.6 × 109 cm−2 was reached after electron propaga-tion in the 1 mm-thick Au-radiator and 2 mm in the activa-tion samples. These values are in a good agreement with the experimental results. We would like to note that the above mentioned ultra-high gamma fluence was generated in the samples situated 180 mm away from the laser-foam interac-tion point. The solid angle covered by the activainterac-tion stack with 2.25 cm2 area is 7× 10−3sr. This means that≤4% of the MeV gamma-rays and the super-ponderomotive electrons generated at the target position participated in the gamma-ray production.

GEANT-4 simulations were performed for an optimized set-up geometry. In simulations, a 1 cm × 1 cm × 0.5 cm Au-block was placed at 1 cm distance from the NCD-target and irradiated by the beam of super-ponderomotive electrons (figure5, 0). In this geometry, 100% of multi-MeV electrons participate in the multi-MeV-bremsstrahlung gen-eration. Table1 shows the number of MeV-photons Nphand effective temperature Teff of the bremsstrahlung spectrum in

(14)

dependence on the photon energy range and the depth in the Au-block. The highest photon number with energies 1– 70 MeV is obtained at 3 mm of the Au-depth and reaches

∼1012 photons. Simulations show that the gamma-beam is directed along the laser axis within a divergence cone of 0.4 sr. The increase of the MeV-photon number with the target thickness can be explained with a long collision free path of multi-MeV-electrons in the target. At a higher thickness (e.g. at 5 mm) the photon attenuation caused by the Compton scat-tering and the pair production starts to play a role and the photon number drops slightly. The effective temperatures of the bremsstrahlung radiation for two photon energy regions presented in table1reflect the two-temperature distribution of the primary electron beam (figure5). The yields of gamma-driven nuclear reactions follow the gamma-fluence trend.

Summarizing the results presented in table1, one can con-clude that interaction of high-current well-directed relativistic electrons with high Z targets leads to effective production of MeV bremsstrahlung radiation with the ultra-high fluence of

∼1012cm−2and the flux of∼1023cm−2s−1assuming a 10 ps gamma-ray pulse in the case of 3 mm target thickness. The conversion efficiency of the laser energy into the energy of gammas beyond >1 MeV reaches 1.5%, while for the region of the giant dipole resonance (GDR) with photon energies E > 7 MeV it is of 0.8%.

6. Summary

The experimental results and numerical simulations demon-strate an extremely high capability of the well-directed high-current relativistic electron beams to be used in novel laser assisted applications exploiting already existing high-energy sub-PW and PW-class laser systems. This makes our approach crucial for applications promising a strong enhancement of the characteristics of the laser driven sources of particles and photons.

In the reported experiment, ultra-relativistic electron beams were produced in the interaction of ∼1019 Wcm−2 laser pulses with plasmas of near critical electron density by the mechanism of direct laser acceleration (DLA). We investig-ated the angular dependence of the electron acceleration pro-cess and measured the high effective electron temperature of 10–13 MeV and the maximum of the electron energy of 100 MeV in the laser pulse propagation direction. A high stability and reproducibility of the acceleration process was observed. The laser energy conversion efficiency into elec-trons with energies above 2 MeV reached 23% with a total charge approaching 1 µC.

For application purposes, we analyzed the charge car-ried by the super-ponderomotive electrons propagating with E > 7.5 MeV. Counting only for electrons in 0.16 sr diver-gence cone, it reaches∼50 nC with the corresponding laser conversion efficiency of∼6%. These results are supported by the full 3D PIC-simulations.

The measured effective electron temperature and the max-imum of the electron energy were twice higher for shots

onto pre-ionized foams at∼1019Wcm−2than for direct laser shots onto standard foils at ultra-relativistic laser intensity of

∼1021Wcm−2. The substantial difference in the electron spec-tra for these two cases presented itself in the isotope produc-tion yield. We detected high yield nuclear reacproduc-tions demand-ing tens of MeV gamma-rays in shots onto pre-ionized foam layers and their combination with foils.

The interaction of the super-ponderomotive electrons with high-Z samples, placed 180 mm away from the target pos-ition, resulted in the photon fluence in the GDR-region of (2.0 ± 0.5) × 109 cm−2 with 12.7 ± 2.4 MeV effective temperature. For the optimized target set-up geometry, we obtain an ultra-high MeV photon number of ∼1012. The gamma-beam is directed along the laser axis within a diver-gence cone of 0.4 sr. For the gamma-ray energies bey-ond 7 MeV, the optimization of the target set-up results in 2 × 1011 photons per laser shot and a corresponding flu-ence of 2 × 1011 cm−2. The effective temperature of the bremsstrahlung spectrum is up to 10 MeV and a correspond-ing laser-to-gamma conversion efficiency of 0.8% obtained at

∼1019Wcm−2laser intensity.

Ultra-intense well-directed beams of MeV electrons and gamma-rays discussed in our paper were obtained at laser intensities that are relevant for the current short pulse high energy diagnostic lasers e.g. at NIF and LMJ. Application of the low-density polymer foams will result in a strong increase of their diagnostic potential in probing of high energy density matter.

Acknowledgments

The results presented here are based on the experiment P176, which was performed at the PHELIX facility at the GSI Helm-holtzzentrum fuer Schwerionenforschung, Darmstadt (Ger-many) in the frame of FAIR Phase-0. The experimental group is very thankful for the support provided by the PHELIX-laser team at GSI-Darmstadt. We thank also Dr B. Borm for the development of the electron spectrometers. The research lead-ing to these results has received fundlead-ing from LASERLAB-EUROPE (grant agreement no. 654148, European Union’s Horizon 2020 research and innovation program). This work was also funded by the Grant Nos. 16 APPA (GSI) of the Min-istry of Science and Higher Education of the Russian Fed-eration, by the Project DFG PU 213/9-1, the RFBR project 19-02-00875, the RFBR and ROSATOM project 20-21-00150, and within the framework of the EUROfusion Consortium by the Euroatom research and training programs 2014−2018 and 2019–2020 under the grant agreement No. 633053. The views and opinions expressed herein do not necessarily reflect those of the European Commission.

ORCID iDs

O N Rosmejhttps://orcid.org/0000-0003-0447-3510 F Consolihttps://orcid.org/0000-0001-8694-3357

(15)

Plasma Phys. Control. Fusion 62 (2020) 115024 O N Rosmej et al

References

[1] Wang T, Ribeyre X, Gong Z, Jansen O, d’Humi`eres E, Stutman D, Toncian T and Arefiev A 2020 Power scaling for collimated g-ray beams generated by structured laser-irradiated targets and its application to two-photon pair production Phys. Rev. Appl.13 054024

[2] Norreys P A et al 1999 Observation of a highly directional g-ray beam from ultrashort, ultraintense laser pulse interactions with solid Phys. Plasmas6 2150

[3] Hatchet S P et al 2000 Electron, photon, and ion beams from the relativistic interaction of Petawatt laser pulses with solid targets Phys. Plasmas7 2076

[4] Zhu X L, Chen M, Weng S-M, Yu T-P, Wang W-M, He F, Sheng Z-M, McKenna P, Jaroszynski D A and Zhang J 2020 Extremely brilliant GeV g-rays from a two-stage laser-plasma accelerator Sci. Adv.6 eaaz7240

[5] Gu Y, Jirka M, Klimo O and Weber S 2019 Gamma photons and electron-positron pairs from ultra-intense laser-matter interaction: A comparative study of proposed configurations

Matter Radiat. Extremes4 064403

[6] Zhu X L, Yu T-P, Sheng Z-M, Yin Y, Turcu I C E and Pukhov A 2016 Dense GeV electron-positron pairs generated by laser in near–critical density plasmas Nat.

Commun.7 13686

[7] Ridgers C P et al 2012 Dense electron-positron plasma and ultraintense g-rays from laser –irradiated solids Phys. Rev.

Lett.108 165006

[8] Pomerantz I et al 2014 Ultrashort pulsed neutron source Phys.

Rev. Lett.113 184801

[9] Kampfrath T, Tanaka K and Nelson K A 2013 Resonant and nonresonant control over matter and light by intense terahertz transients Nat. Photonics7 680

[10] Koenig S et al 2013 Wireless sub-THz communication system with high data rate Nat. Photonics7 977

[11] Herzer S et al 2018 An investigation on THz yield from laser-produced solid density plasmas at relativistic laser intensities New J. Phys.20 063019

[12] Kostyukov I, Kiselev S and Pukhov A 2003 X-ray generation in an ion channel Phys. Plasmas10 4818–28

[13] Kislev S, Pukhov A and Kostyukov I 2004 X-ray generation in strongly nonlinear plasma wave Phys. Rev. Lett.93 135004

[14] Rousse A et al 2004 Production of a keV X-ray beam from synchrotron radiation in relativistic laser-plasma interaction

Phys. Rev. Lett.93 135004

[15] Cipiccia S et al 2011 Gamma-rays from harmonically resonant betatron oscillation in plasma wake Nat. Phys. Lett.

7 867–71

[16] Albert F et al 2017 Observation of betatron x-ray radiation in a self-modulated laser wakefield accelerator driven with picosecond laser pulses Phys. Rev. Lett.118 134801

[17] Kneip S et al 2008 Observation of synchrotron radiation from electrons accelerated in a petawatt-laser-generated plasma cavity Phys. Rev. Lett.100 105006

[18] Ravasio A et al 2008 Hard x-ray radiography for density measurement in shock compressed matter Phys. Plasmas

15 060701

[19] Li K, Borm B, Hug F, Khaghani D, Löher B, Savran D, Tahir N A and Neumayer P 2014 Developments toward hard X-ray radiography on heavy-ion heated dense plasmas

Laser Part. Beams32 631

[20] Negoita F et al 2016 Laser driven nuclear physics at ELI-NP

Rom. Rep. Phys. 68 37–144

[21] Habs D and Köster U 2011 Production of medical radioisotopes with high specific activity in photonuclear reactions with γ-beams of high intensity and large brilliance Appl. Phys. B103 501–19

[22] Ma Z, Lan H, Liu W, Wu S, Xu Y, Zhu Z and Luo W 2019 Photonuclear production of medical isotopes 62,64Cu using

intense laser-plasma electron source Matter Radiat.

Extremes4 064401

[23] Esarey E, Schroeder C B and Leemans W P 2009 Physics of laser-driven plasma-based electron accelerators Rev. Mod.

Phys.81 1229

[24] Pukhov A and Meyer-ter-vehn J 2001 Laser wake field acceleration: the highly non-linear broken-wave regime

Appl. Phys. B74 355

[25] Faure J, Glinec Y, Pukhov A, Kiselev S, Gordienko S, Lefebvre E, Rousseau J-P, Burgy F and Malka V 2004 A laser-plasma accelerator producing monoenergetic electron beams Nature431 541

[26] Mangles S et al 2004 Monoenergetic beams of relativistic electrons from intense laser–plasma interactions Nature

431 535

[27] Geddes C G R, Toth C, van Tilborg J, Esarey E, Schroeder C B, Bruhwiler D, Nieter C, Cary J and Leemans W P 2004 High-quality electron beams from a laser wakefield accelerator using plasma-channel guiding

Nature431 538

[28] Gonsalves A J et al 2019 Petawatt laser Guiding and electron beam acceleration to 8 GeV in a laser-heated capillary discharge waveguide Phys. Rev. Lett.122 084801

[29] Huang Z, Ding Y and Schroeder C B 2012 Compact x-ray free-electron laser from a laser-plasma accelerator using a transverse-gradient undulator Phys. Rev. Lett.

109 204801

[30] Maier A R et al 2012 Demonstration scheme for a laser-plasma-driven free-electron laser Phys. Rev. X 2 031019

[31] Seggebrock T, Maier A R, Dornmair I and Grüner F 2013 Bunch decompression for laser-plasma driven free-electron laser demonstration schemes Phys. Rev. Spec. Top. Accel

Beams16 070703

[32] Pukhov A, Sheng Z-M and Meyer-ter-vehn J 1999 Particle acceleration in relativistic laser channels Phys. Plasmas

6 2847

[33] Pukhov A 2003 Strong field interaction of laser radiation Rep.

Prog. Phys.66 47–101

[34] Arefiev A V, Khudik V N, Robinson A P L, Shvets G, Willingale L and Schollmeier M 2016 Beyond the ponderomotive limit: direct laser acceleration of relativistic electrons in sub-critical plasmas Phys. Plasmas

23 056704

[35] Willingale L et al 2018 The unexpected role of evolving longitudinal electric fields in generating energetic electrons in relativistically transparent plasmas New J. Phys.

20 093024

[36] Bochkarev S G, Brantov A V, Bychenkov V Y, Torshin D V, Kovalev V F, Baidin G V and Lykov V A 2014 Stochastic electron acceleration in plasma waves driven by a high-power subpicosecond laser pulse Plasma Phys. Rep.

40 202–14

[37] Bagnoud V et al 2010 Commissioning and early experi-ments of the PHELIX facility Appl. Phys. B

100 137–50

[38] Gray R J et al 2014 Laser pulse propagation and enhanced energy coupling to fast electrons in dense plasma gradient

New J. Phys.16 113075

[39] Willingale L et al 2013 Surface waves and electron acceleration from high-power, kilojoule-class laser interaction with underdense plasma New J. Phys.

15 025023

[40] Pugachev L P et al 2016 Acceleration of electrons under the action of petawatt-class laser pulses onto foam targets Nucl.

Instrum. Methods Phys. Res. A829 88–93

[41] Pugachev L P and Andreev N E 2019 Characterization of accelerated electrons generated in foams under the action of petawatt lasers J. Phys.: Conf. Ser. 1147 012080

(16)

[42] Borisenko N G, Khalenkov A M, Kmetik V, Limpouch J, Merkuliev Y A and Pimenov V G 2007 Plastic aerogel targets and optical transparency of undercritical microheterogeneous plasma Fusion Sci. Technol.

51 655–64

[43] Rosmej O N et al 2019 Interaction of relativistically intense laser pulses with long-scale near critical plasmas for optimization of laser based sources of MeV electrons and gamma-rays New J. Phys.21 043044

[44] Gus’kov S Y, Limpouch J, Nicolaï P and Tikhonchuk V T 2011 Laser-supported ionization wave in under-dense gases and foams Phys. Plasmas18 103114

[45] Nicolaï P et al 2012 Experimental evidence of foam homogenization Phys. Plasmas19 113105

[46] Khalenkov A M, Borisenko N G, Kondrashov V N, Merkuliev Y A, Limpouch J and Pimenov V G 2006 Experience of micro-heterogeneous target fabrication to study energy transport in plasma near critical density Laser

Part. Beams24 283–90

[47] Borisenko N G et al 2006 Regular 3-D networks with clusters for controlled energy transport studies in laser plasma near critical density Fusion Sci. Technol.49 676–85

[48] Bonnet T, Comet M, Denis-Petit D, Gobet F, Hannachi F, Tarisien M, Versteegen M and Al´eonard M M 2013 Response functions of imaging plates to photons, electrons and 4He particles Rev. Sci. Instr.84 103510

[49] Tanaka K A, Yabuuchi T, Sato T, Kodama R, Kitagawa Y, Takahashi T, Ikeda T, Honda Y and Okuda S 2005 Calibration of imaging plate for high energy electron spectrometer Rev. Sci. Instr.76 013507

[50] Rusby D R et al 2015 Measurement of the angle, temperature and flux of fast electrons emitted from intense laser-solid interactions J. Plasma Phys.81 475810505

[51] Stoyer M A et al 2001 Nuclear diagnostics for petawatt experiments Rev. Sci. Instr.72 767–72

[52] Günther M M, Sonnabend K, Brambrink E, Vogt K, Bagnoud V, Harres K and Roth M 2011 A novel nuclear pyrometry for the characterization of the high-energy bremsstrahlung and electrons produced in relativistic laser-plasma interactions Phys. Plasmas18 083102

[53] Zankl G, Strachan J D, Lewis R, Pettus W and Schmotzer J 1981 Neutron flux measurements around the Princeton large tokamak Nucl. Instrum. Methods Phys. Res.

185 321–9

[54] Bin J et al 2018 Enhanced laser-driven ion acceleration by superponderomotive electrons generated from

near-critical-density plasma Phys. Rev. Lett.120 074801

[55] Pukhov A 1999 Tree-dimensional electromagnetic relativistic particle-in-cell code VLPL (virtual laser plasma lab)

Plasma Phys.61 425–33

[56] Geant4 Toolkit (available at:http://geant4-userdoc.web. cern.ch/geant4-userdoc/UsersGuides/ForApplication Developer/fo/BookForApplicationDevelopers.pdf) [57] Geant4 Toolkit, physicslist (available at:

http://geant4-userdoc.web.cern.ch/geant4-userdoc/UsersGuides/ PhysicsListGuide/fo/PhysicsListGuide.pdf)

[58] GEANT4 Physics Reference Manual Release 10.6 p 401 [59] Koning A J and Rochman D 2012 Modern nuclear data

evaluation with the TALYS code system Nucl. Data Sheets

Figura

Figure 1. Experimental set-ups (top-view) used for shots (a) at moderate and (b) ultra-high relativistic intensities
Figure 2. Raw electron spectra measured along the laser axis by means of the 0.99 T electron spectrometer (a, c) and angular distribution of the electron beam with E > 3 MeV and > 7.5 MeV registered using the cylinder-stack (b, d): (a, b) shot at 1.6
Figure 4. Energy spectra of electrons per steradian measured at 0 ◦ (red), 15 ◦ (green) and 45 ◦ (blue) to the laser propagation direction for a shot onto the pre-ionized foam layer at 1.9 × 10 19 Wcm − 2 laser intensity.
Figure 6. Electron energy distributions are normalized to the corresponding laser energies inside FWHM of the focal spot in Joules
+6

Riferimenti

Documenti correlati

Finally, the results presented in this paper can be also considered as an additional theoretical confirmation that the use of SAW to drive electrons along quantum wires prevents

La campagna di scavo, ancora in corso, ha portato alla luce i resti scheletrici di ben 400 individui, verosimilmente deceduti durante l’epidemia di peste che devastò la città

technique) with ground multi- and hyperspectral spectral observations, as well as quantitative estimation of vegetation biophysical variables from proximal sensing.

il Vero, cit., pp. La testa di Vittoria par- tecipò a Napoli alla mostra alla Galleria Chiaia di Napoli nel ’47 e alla II Esposi- zione di Arti figurative del ’48; in seguito, alla II

The evidence of micro-scale heterogeneities in the products of the crystal-poor evolved units (from PPF to BU/SU), suggests that several evolved melts were present at the

Proprio a partire dall’uso del paradigma dell’Esodo è perciò possibile rileg- gere e classificare le opere appartenenti alla letteratura sulla Respublica Hebra- eorum, la cui

The main result is that, if all the deformations of the satellite dissipate some energy, then under a suitable nondegeneracy condition there are only three possible outcomes for

Indeed, the maps make evident the presence of inhomogeneities, but do not provide quantitative details on the tunnelling conductance values; on the other hand, the dI /dV line